You are on page 1of 12

Spectrochimica Acta Part A 60 (2004) 689–700

Lattice vibrations of manganese oxides


Part I. Periodic structures
C.M. Julien a,∗ , M. Massot b , C. Poinsignon c
a Laboratoire des Milieux Désordonnés et Hétérogènes, CNRS-UMR 7603, Université Pierre et Marie Curie,
4 Place Jussieu, Case 86, 75252 Paris Cedex 05, France
b Laboratoire de Physique des Milieux Condensés, CNRS-UMR 7602, Université Pierre et Marie Curie,

4 Place Jussieu, Case 77, 75252 Paris Cedex 05, France


c Laboratoire d’Electrochimie, de Physicochimie des Matériaux et des Interfaces, CNRS INPG, UMR 5631,

ENSEEG, Domaine Universitaire BP 75, 38402 Saint Martin d’Hères, France

Received 13 June 2003; received in revised form 13 June 2003; accepted 28 June 2003

Abstract

Raman scattering (RS) and Fourier transform–infrared (FT–IR) spectroscopy have been applied to the structural characterisation of man-
ganese dioxides (MDOs). A variety of synthetic battery-grade MDOs are investigated for comparison to the natural phases. The RS and
FT–IR spectra are analysed on the basis of the local environment in the MDO structures considering the vibrations of the MnO6 octahedral
building the lattices. The vibrational modes of the MnO6 units expand over 400–650 cm−l with additional bands in the low-wavelength region.
Structural trends are deduced from the comparison of the vibrational spectra of the MDO phases investigated: birnessite, bixbyite, coronadite,
groutite, hausmannite, hollandite, manganosite, pyrolusite, ramsdellite, romanechite, spinel, and todorokite.
© 2003 Elsevier B.V. All rights reserved.

Keywords: Manganese oxides; Lattice dynamics; Raman spectroscopy; FT–IR spectroscopy; Lithium batteries

1. Introduction There is a wide variety of battery-grade manganese


dioxides (MDOs), whose structures have been extensively
Manganese oxides with tunnel and layered crystal struc- investigated [6]. Among the MDOs, most attention has been
tures constitute a large family of porous materials of differ- focused on synthetic products prepared by either electrolytic
ent pore size [1] largely used as catalysts, oxidant agents, (EMD) or chemical (CMD) method that belong to the nsutite
absorbent and electrode materials. In the field of energy (␥-MnO2 ) group for their use as cathode material in alkaline
storage, the excellent electrochemical properties of sev- batteries [7]. Besides the ␥-MnO2 form, particular attention
eral manganese-oxide (MO) phases are attracting much has been paid to stoichiometric compounds. As an exam-
attention for positive electrode materials for lithium and ple, the sodium-free birnessite, MnO1.85 ·0.6H2 O, displays a
alkaline batteries [2–4]. MnO2 was originally developed layered structure with larger interlayer distance and trigonal
as the positive electrode for primary Li cell. Extensive re- prismatic sites favourable for easy lithium insertion [8].
searches were carried out during the last decades to improve Raman scattering (RS) and Fourier transform–infrared
the reversibility of lithium insertion in MO cathode for (FT–IR) spectroscopy have proved valuable for the char-
rechargeable Li//MnO2 cells [5]. This application generates acterisation of lithiated transition-metal oxides used as
an important activity in the field of Solid State Chemistry electrode materials in rechargeable lithium batteries [9–20].
to synthesise new phases of MO with improved electro- Bulk and powder manganese oxides were analysed using
chemical properties as stability during cycling, with large Raman [21–30] and infrared spectroscopy [31–41]. A care-
reversible capacity under fast discharge rate. ful examination of the results published over 40 years shows
discordance regarding the Raman spectra of oxides. This is
∗ Corresponding author. Tel.: +33-1-44-27-4561; mainly due to the poor structural characterisation of MDOs
fax: +33-1-44-27-4512. of complex crystal chemistry whom structures request a
E-mail address: cjul@ccr.jussieu.fr (C.M. Julien). careful analysis. This is due also to the great unstability of

1386-1425/$ – see front matter © 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S1386-1425(03)00279-8
690 C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700

Table 1 ture are listed in Table 1. More than 35 MnO2 samples


Characteristics of the synthetic and mineralogical varieties of manganese have been examined critically. Mineralogical manganese
oxides
dioxide and non-dioxides samples include: orthorhombic
Sample Designation Origin Structure ramsdellite (from Wiskenburg, Arizona), (ii) monoclinic
A Ramsdellite Wiskenburg, Arizona Orthorhombic romanechite (from Romanèche, France), (iii) tetragonal
B ␤-MnO2 pyrolusite Aldrich Rutile hollandite (from Jabuha, India), (iv) tetragonal corona-
D ␣-MnO2 Synthetic Tetragonal dite (from Imini, Morocco), (v) orthorhombic groutite
E Hollandite Jabuha, India Tetragonal
F Todorokite Japan Monoclinic
(from Cuyana Range, Minnesota), (vi) monoclinic todor-
G Romanechite Romanèche, France Monoclinic okite (from Japan), and (vii) pseudo-orthorhombic nsu-
H Coronadite Imini, Morocco Tetragonal tite or natural manganese dioxide (NMD) from Ghana.
I Birnessite Sol–gel synthesis Hexagonal ␭-MnO2 form is an anhydrous product prepared by chemi-
J ␭-MnO2 spinel Synthetic Cubic cal delithiation of ␭-LiMn2 O4 in acid as reported by Hunter
K ␣-Mn2 O3 bixbyite Alfa Aesar Cubic
L ␥-Mn2 O3 Synthetic Tetragonal
[3].
M Mn3 O4 hausmannite Alfa Aesar Tetragonal The structure of the studied samples was characterised
N MnO manganosite Alfa Aesar Cubic by X-ray powder diffraction (XRPD) using a diffractome-
O MnOOH groutite Cuyana Range, Orthorhombic ter (Philips model PW1830) with nickel-filtered Cu K␣
Minnesota radiation (λ = 1.5406 Å). The diffraction patterns were
taken at room temperature in the range of 5◦ < 2θ < 80◦
MDOs, under the laser beam; most of them can be reduced using step scans. Raman scattering spectra were taken
to the stable phase Mn2 O3 (bixbyite). The formation of between 10 and 1200 cm−1 at room temperature in a
spinel zones at the laser impact point is evidenced by the quasi-backscattering configuration using a Raman-laser
appearance of a strong RS peak at 650 cm−1 [20]. More bench apparatus equipped with holographic grating dual
reproducible results were obtained within infrared results. monochromator and intermediate spatial filter (Jobin-Yvon
In this respect, the excellent work of Potter and Rossman U1000 model). The laser light source was the emission
must be referenced: it provides a series of infrared powder of the argon-ion laser (Spectra-Physics 2020 model) at
absorption spectra of most naturally occurring tetravalent wavelength 514.5 nm. Each RS spectrum is the average
and trivalent manganese oxides [41]. of 12 successive scans obtained at a spectral resolution
The lattice dynamics of lithium manganese oxides hav- of 2 cm−1 to gain a high signal/noise ratio. Frequency
ing spinel-related structure were presented previously stability and accuracy of the apparatus were checked in
[9–12,20]. The vibrational spectra of the LiMn2 O4 were recording the Raman spectrum of silicon. To avoid sam-
studied using RS and FT–IR techniques, which probe di- ple photo transformation RS spectra were recorded us-
rectly the near-neighbouring environment of oxygen coor- ing a low excitation power of 10 mW (power density of
dination around the lithium and manganese cations. Using 100 W cm−2 ). An increase in lattice temperature generally
the spectroscopic symmetry O7h , the normal modes of the results in a shift of Raman peak wavenumber up to the
spinel LiMn2 O4 are discussed with the identification of the formation of Mn2 O3 oxide. All commercial MO powders
vibrations due to the Mn–O and Li–O polyhedra. were pelletised to obtain a mirror-like surface sample for
In this paper, vibrational spectra of several manganese Raman analysis. Infrared absorption spectra were recorded
oxides are analysed using RS and FT–IR spectroscopy at room temperature using a Fourier transform interfer-
as complementary techniques in the wavenumber ranging ometer (model Bruker IFS113v). This vacuum bench ap-
100–1000 cm−1 to investigate internal and external modes paratus was equipped with a Mylar 3.5 ␮m-thick beam
of MDO lattices. The samples include periodic structures, splitter, a globar source, and a DTGS/PE far-infrared de-
such as birnessite, bixbyite, coronadite, groutite, hausman- tector. Samples were ground to fine powders dispersed
nite, hollandite, manganosite, pyrolusite, ramsdellite, ro- into ICs pellet or coated over solid-paraffin slab which
manechite, spinel, and todorokite, whose characteristics are are non-absorbing media in the investigated wavenum-
presented in Table 1. Our ambition is to establish a com- ber range. Data were collected in transmission mode at a
parative spectral identification of these different manganese spectral resolution of 2 cm−1 after 256 scans in vacuum
oxides. Therefore, our analysis is restricted to the group atmosphere.
frequency approximation. The nsutite family, represented Raman and FT–IR spectra were deconvoluted using
by the disordered synthetic ␥-MnO2 , will be considered in the curve analysis based on the original algorithm of
the second part of this series. non-linear peak fitting known as the Levenberg–Marquardt
method. The deconvolution analysis of IR spectra was
done assuming a linear baseline in the spectral range of
2. Experimental 50–1200 cm−1 . All the Raman bands introduced in the fit
had a Lorentzian lineshape; two adjusting parameters (inten-
Measurements were performed on synthetic and min- sity and band-width) were considered for each Lorentzian
eralogical varieties of MnO2 , whose reference and struc- oscillator.
C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700 691

3. Structural properties of manganese dioxides to the number of MnO6 units and the number of MnO6
octahedral chains between two basal layers to form tunnel
Manganese ion presents three different valence states (Tm,n ) openings. The T1,n group includes two chemically
in the series of studied oxides of complex structure pure forms: the pyrolusite ␤-MnO2 (T1,1 ) and the ramsdel-
illustrated by the numerous forms of mineralogical lite R–MnO2 (T1,2 ). The Tm,∞ group includes the layered
manganese dioxides: tetragonal pyrolusite ␤-MnO2 , or- phyllomanganates such as birnessite, buserite, and rancieite.
thorhombic ramsdellite, hollandite group (K,Pb)Mn8 O16 , Fig. 1 shows the schematic representation of the various
psilomelane group (Ba,H2 O)2 Mn5 O10 , todorokite group manganese dioxide frameworks showing the variation in the
(Mn,Ca,Mg)Mn3 O7 ·H2 O, etc. All MnO2 structures can be chain and tunnel (m × n) structures: (a) pyrolusite (T1,1 ),
described as a close-packed network of oxygen atoms in (b) ramsdellite (T1,2 ), (c) hollandite (T2,2 ), (d) birnessite
which Mn4+ cations (ionic radius of 0.53 Å) are differently (T1,∞ ), (e) romanechite (T2,3 ), and (f) spinel (T1,1 ).
distributed. These distributions are conveniently described The stable form of MnO2 is the mineral pyrolusite or
by filled [MnO6 ] octahedra sharing opposite octahedral ␤-MnO2 . It has the tetragonal rutile structure (P42 /mnm
edges to form endless chains, which can in turn be linked space group) containing single chains running along the
to neighbouring octahedral chains by sharing corners or c-axis, sharing only corners with adjacent chains. All oc-
edges. The manganese dioxides can be classified according tahedra are equivalent but distorted with an average Mn–O

Fig. 1. Schematic representation of the various manganese dioxide frameworks showing the variation in the chain and tunnel (m × n) structures: (a)
pyrolusite (1 × 1), (b) ramsdellite (1 × 2), (c) hollandite (2 × 2), (d) birnessite (1 × 8), (e) romanechite (2 × 3), and (f) spinel (1 × 1).
692 C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700

Table 2
Crystallographic data of some Mn–O compounds
Compound Mineral Crystal symmetry Lattice parameters (Å) Features

MnO Manganosite Cubic (Fm3m) a = 4.44 Rock-salt


␣-MnO2 Hollandite Tetragonal (I4/m) a = 9.96; c = 2.85 (2 × 2) Tunnel
R–MnO2 Ramsdellite Orthorhombic (Pbnm) a = 4.53; b = 9.27; c = 2.87 (1 × 2) Tunnel
␤-MnO2 Pyrolusite Tetragonal (P42 /mnm) a = 4.39; c = 2.87 (1 × 1) Tunnel
␥-MnO2 Nsutite Complex tunnel (hex) a = 9.65; c = 4.43 (1 × 1)/(1 × 2)
␦-MnO2 Vernadite Hexagonal a = 2.86; c = 4.7 (1 × ∞) Layer
␭-MnO2 Spinel Cubic (Fd3m) a = 8.04 (1 × 1) Tunnel
MnOx ·H2 O Birnessite Tetragonal ahex = 2.84; chex = 14.64 (1 × ∞) Layer
MnOOH Groutite Orthorhombic (Pbnm) a = 4.56; b = 10.70; c = 2.87
␣-Mn2 O3 Bixbyite Cubic (Ia3) a = 9.41 C-type
Mn3 O4 Hausmannite Tetragonal (I41 /amd) a = 9.81; c = 2.85 Spinel-like

distance of 1.86 Å. However, the bridging Mn–O distance Fig. 2 shows the XRPD pattern of natural and com-
within a chain is shorter than the apical Mn–O distance mercial manganese dioxides, i.e. ramsdellite R–MnO2 ,
within the basal planes. The structure consists of strings ␤-MnO2 , synthetic ␣-MnO2 , spinel ␭-MnO2 , birnes-
of MnO6 octahedra and empty channels corresponding to site MnO1.86 ·0.6H2 O, bixbyite Mn2 O3 , and hausmannite
a width of (1 × 1) octahedron. The orthorhombic ramsdel- Mn3 O4 . X-ray diffraction pattern of ␭-MnO2 displays the
lite (Pbnm space group) contains double chains formed by [Mn2 ]O4 spinel framework structure (Fig. 1(f)), where
edge-sharing [MnO6 ] octahedra. It is a metastable form of the Mn cations are distributed in alternate layers in a
MnO2 containing (1 × 2) blocks of MnO6 octahedra and 3:1 ratio. The interstitial space of ␭-MnO2 contains a
empty channels of the same size. Another tetragonal form three-dimensional network of interlinked (1 × 1) chan-
␣-MnO2 (I4/m space group) contains (2×2) open tunnels in nels. The crystallographic structure of ␭-MnO2 possesses
a network of MnO6 octahedra [1]. Crystallographic data re- the symmetry Fd3m and has a general structural formula
garding the various structures of MO samples are presented A[Mn2 ]O4 , where the Mn4+ cations reside on the octahe-
in Table 2. dral (16d) sites, the oxygen anions occupy the (32e) sites,
and A are voids.
The XRD pattern of birnessite-type MnO2 powders dis-
(a) plays at low angles, i.e. around 2θ = 12◦ , the intense
line due to the basal spacing of the lamellar structure.
The structure of MnO1.85 ·0.6H2 O is trigonal (P 3̄m1 space
group) with the unit cell parameters a = b = 2.85 Å and
(b)
c = 7.25 Å. The preferential orientation shown by the
strong appearance of the (0 0 1) and (0 0 2) Bragg lines cor-
responds to the alignment of MnO6 layers parallel to the (a,
XRD intensity (a. u.)

(c) b) planes. The manganese sesquioxide Mn2 O3 (bixbyite)


has the C-type structure. This structure is body-centred cu-
bic, space group Ia3 (Th7 ) which contains eight Mn3+ ions
(d) located on the (b) sites with symmetry S6 and 24 Mn3+
on the (d) sites with point symmetry C2 . The 48 oxygens
in the body-centred cell are on general positions (e) with
(e) C1 site symmetry. The hausmannite Mn3 O4 (MnMn2 O4 in
spinel notation) crystallises in the normal tetragonal spinel
structure (A[B2 ]O4 type) with space group I41 /amd.
(f)

4. Lattice vibrations of T1,n manganese dioxides


(g)
Due to their electronic properties, the general peculiarity
10 20 30 40 50 60 70 80 of the vibrational features of MOs is their low Raman activ-
2 theta (degree) ity. Three major regions, in which the band frequencies shift
as a function the local structure, can be recognised in the
Fig. 2. X-ray powder diffraction patterns of various manganese oxides: (a)
ramsdellite R-MnO2 , (b) ␤-MnO2 (from Aldrich), (c) synthetic ␣-MnO2 , Raman and FT–IR spectra of MOs; they appear at 200–450,
(d) spinel ␭-MnO2 , (e) birnessite MnO1.86 ·0.6H2 O, (f) bixbyite Mn2 O3 , 450–600, and 600–750 cm−1 . They correspond to spectral
and (g) hausmannite Mn3 O4 (IBA100 sample). domains where wagging, bending and stretching vibrations
C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700 693

Table 3 100
Raman wavenumber of vibrational modes for various manganese oxides
Compound Raman wavenumber (cm−1 )

538
␯1 ␯2 ␯2 ␯4 ␯5 ␯5 ␯6 ␯7 ␯8

665
␤-MnO2 – 319 377 486 538 – 665 – 750 75
486

␥-MnO2 a
319

377

264 337 379 491 520 572 631 670 738

750
RS intensity (cps)

R–MnO2 275 – 387 490 522 575 630 648 742 (b)
␣-Mn2 O3 192 314 404 481 – 592 645 698 –
␥-Mn2 O3 263 308 – – 512 – 631 670 –
Mn3 O4 – 310 357 485 579 – 653 – 50
575

MnO 250 – – 531 591 – 654 – –


MnOOH 213 253 352 384 528 552 615 648 –
522

a With pyrolusite intergrowth Pr = 29%.

25
387

648

are active modes, respectively. Band frequencies determined


275

490

630

on the RS and FT–IR spectra of manganese oxides are pre-


742

(a)
sented in Tables 3 and 4, respectively.
Few works reported the vibrational spectra of ␤-MnO2
and ␥-MnO2 lattices. A careful examination of the spectra 0
200 300 400 500 600 700 800 900
published shows discordance regarding RS spectra due to the
poor structural characterisation of MDO networks [21–26]. Wavenumber (cm -1)
Figs. 3 and 4 show the Raman scattering and FT–IR spec- Fig. 3. Raman scattering spectra of: (a) ramsdellite R–MnO2 and (b)
tra of the end members of T1,n MDO compounds, i.e. pyro- pyrolusite ␤-MnO2 samples.
lusite ␤-MnO2 and ramsdellite R–MnO2 , respectively. As an
experimental fact, the Raman signal of ␤-MnO2 is weaker
than that of ramdellite due to the quasi metallic state of the the skeletal vibrations. The main contributions are attributed
former material for which the laser beam penetration depth to the stretching mode of the Mn–O bond in MnO6 octahe-
is very low. dra, the lower frequency peaks to the deformation modes of
Rutile-type MnO2 belongs to the D14 the metal-oxygen chain of Mn–O–Mn in the MnO2 octahe-
4h spectroscopic group
(Z = 2) with Raman activity such as 1A1g + 1B1g + 1B2g + dral lattice. The corresponding asymmetric stretching modes
1Eg . In this structure, the oxygen ions triply bridge three are recorded in the FT–IR spectrum at 517 and 621 cm−1 .
Mn ions and lie in a C2v symmetry site. So, three modes The infrared features of ␤-MnO2 (Fig. 4, curve b) dis-
are expected with A1 , B1 and B2 symmetries. B1 modes are play five bands which are in good agreement with the
definitely out-of-plane deformation modes, while A1 and B2 predictions of the factor group analysis for the tetragonal
modes mix the character of Mn–O stretching with those of
Mn–O in-plane deformation modes [29].
2.0
The RS spectrum of ␤-MnO2 (Fig. 3, curve b) has rela-
545

tively sharp peaks indicative of a well-developed rutile-type


618

MnO2 structure with an interstitial space consisting of nar-


row one-dimensional (1 × 1) channels. It is characterised by 1.5
two sharp peaks at 538 and 665 cm−1 along with four weak
387

bands at 319, 377, 486, and 750 cm−1 which correspond to


Absorbance units

329

626

(b)
515

Table 4 1.0
471

Infrared wavenumber of vibrational modes for various manganese oxides


IR wavenumber (cm−1 )
589

Compound
687

␯1 ␯2 ␯3 ␯4 ␯5 ␯6 ␯7 ␯8 ␯9
366

740

0.5
␤-MnO2 – 329 387 – – 545 618 626 –
263

␥-MnO2 a – – 374 480 530 564 – 705 – (a)


R–MnO2 263 – 366 471 515 589 – 687 740
␣-Mn2 O3 – 300 408 453 533 576 602 672 –
␥-Mn2 O3 217 295 390 446 484 523 591 666 – 0.0
200 300 400 500 600 700 800
Mn3 O4 122 164 241 291 356 475 527 594 655
MnO 258 329 – 473 – – – – – Wavenumber (cm -1)
MnOOH 296 329 360 380 510 580 616 995 1022
Fig. 4. FT–IR powder absorption spectra of: (a) ramsdellite R-MnO2 and
a With pyrolusite intergrowth Pr = 29%. (b) pyrolusite ␤-MnO2 samples.
694 C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700

structure, i.e. A2u + 3Eu infrared-active modes. The FT–IR 60


pattern of ␤-MnO2 has relatively sharp peaks indicative of a (a) hollandite
(b) coronadite
well-developed rutile-type structure with an interstitial space (c) romanechite
consisting of narrow one-dimensional (1 × 1) channels. The 50
(d) todorokite
(e) birnessite
(a)
dominant band at 545 cm−1 has been attributed to the vibra-
RS intensity (arb. units)

tion due to the displacement of the oxygen anions relative


to the manganese ions along the direction of the octahedral 40
chains. This mode possesses the A2u symmetry. Two bands (b)
are recorded at higher wavenumebers: at 618 and 626 cm−1 .
Moreover, the specific fingerprints of the vibrational features 30
of ␤-MnO2 are the well-defined and intense far-infrared
bands at 329 and 387 cm−1 , which are attributed to vibra- (c)
tions perpendicular to the octahedral chains (Eu symmetry). 20
The ramsdellite MnO2 structure belongs to the D16 2h
(d)
spectroscopic group (Z = 4) with Raman activity such as
6A1g + 3B1g + 6B2g + 3B3g [31]. The RS spectrum of 10
R-MnO2 (Fig. 3, curve a) displays three main contributions
at 518–520, 575–580 and 630 cm−1 along with three weak (e)
peaks recorded at 275, 387 and 742 cm−1 . The two sharp
0
high-frequency peaks at 648 and 742 cm−1 are indicative of 200 300 400 500 600 700 800 900
a well-developed orthorhombic structure with an interstitial Raman shift (cm -1)
space consisting of (1 × 2) channels. There are assigned to
the A1g spectroscopic species. Furthermore, it is obvious Fig. 5. Raman scattering spectra of manganese dioxides with various
tunnel structures: (a) hollandite ␣-MnO2 , (b) natural romanechite, (c)
that the Raman spectrum of R–MnO2 exhibits features dif-
natural coronadite, (d) natural todorokite, and (e) synthetic birnessite
ferent from those of ␤-MnO2 . This presumably implies that MnO1.86 ·0.6H2 O.
the anharmonicity present in the Mn–O bond of octahedra
in ramsdellite is less than rutile. According to the above
results, the stretching vibrations of MnO6 octahedra lie in ␦-MnO2 (T1,∞ ), and the spinel structure ␭-MnO2 (T1,1 )
the region of 500–700 cm−1 . made of a three-dimensional network of interlinked chan-
The FT–IR absorption spectrum of R–MnO2 displays nels. Figs. 5 and 6 show the Raman scattering and FT–IR ab-
seven bands in the range of 250–750 cm−1 (Fig. 4, curve sorption spectra, respectively, of these manganese dioxides
a). Factor group analysis predicts 12 IR-active modes with
5B1u +2B2u +5B3u symmetry for R–MnO2 [31]. The FT–IR
1.4
patterns show two strong bands in the high-wavenumber re-
(a) hollandite
gion at 740 and 687 cm−1 , which are the main fingerprints (b) coronadite
(c) romanechite
of the ramsdellite structure. The low-wavenumber bands at 1.2
(d) todorokite
283 and 366 cm−1 are very weak and show a larger split- (e) birnessite (a)
Absorbance (arb. units)

ting compared with that of pyrolusite. The dominant absorp- 1.0


tion band at 515 cm−1 is assigned to the vibration due to
the displacement of the oxygen anions relative to the man-
ganese ions along the direction of the octahedral chains. 0.8
Our spectroscopic data on T1,n manganese dioxides are in (b)
good agreement with previous results [32,42,43]. However, 0.6
the mixed T1,1 –T1,2 structures (aperiodic manganese diox-
ides) such as ␥-MnO2 are not considered here, but will be 0.4
(c)
investigated in the following paper of this series.
(d)
0.2
5. Lattice vibrations of Tm,n manganese dioxides (e)
0.0
The manganese dioxides can be classified according the 100 200 300 400 500 600 700 800 900
nature of the organisation of MnO6 units and the number of Wavenumber (cm-1)
MnO6 octahedral chains between two basal layers to form
Fig. 6. FT–IR powder absorption spectra of dioxides with various
tunnel (Tm,n ) openings. They include the hollandite-type tunnel structures: (a) hollandite ␣-MnO2 , (b) natural romanechite, (c)
structure ␣-MnO2 (T2,2 ), the coronadite (T2,2 ), the romane- natural coronadite, (d) natural todorokite, and (e) synthetic birnessite
chite (T2,3 ), the todorokite (T3,3 ), the birnessite layered MnO1.86 ·0.6H2 O.
C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700 695

Table 5 Two minerals with large opening channels were stud-


Raman modes of mineral manganese dioxides ied: romanechite T2,3 , and todorokite T3,3 . They have a
Compound Raman wavenumber (cm−1 ) monoclinic symmetry and present broad channels generated
␯1 ␯2 ␯3 ␯4 ␯5 ␯6 ␯7
by the triple chains of octahedra which respectively con-
tain Ba2+ and water, Mg2+ , Co2+ and/or Zn2+ . The RS
Nsutite 280 382 458 515 572 634 732 spectrum of romanechite displays two strong bands at 643
Hollandite 259 410 507 586 628 –
Coronadite – 332 388 495 585 626 –
and 578 cm−1 and four weak bands at 284, 372, 515 and
Romanechite 284 – 372 515 578 643 721 721 cm−1 , while the RS spectrum of todorokite exhibits
Todorokite 263 – 358 – 590 641 – one dominant band at 641 cm−1 and three weak compo-
Birnessite 296 303 – 506 575 656 730 nents at 263, 358, and 590 cm−1 . The same major IR band
at 520 ± 10 cm−1 is present in romanechite as in nsutite,
ramsdellite, and the hollandite group spectra, with differ-
except the one’s of ␭-MnO2 . RS and FT–IR frequencies of ence in relative intensity. This is attributed to their common
NMD samples are presented in Tables 5 and 6, respectively. crystal chemistry based on edge-shared MnO6 octahedra,
The structure of the hollandite-type compounds (Fig. 1(c)) which provides the ramsdellite-like ␯1 and ␯2 modes located
consists of double chains of edge sharing MnO6 octahedra. at 748 and 716 cm−1 in the FT–IR spectrum of romane-
The term hollandite is used to classify these materials with chite ((Ba,H2 O)2 Mn5 O10 ). Far-infrared bands at 275 and
T2,2 channels but also defines a specific manganese oxide 319 cm−1 are diagnostic of the romanechite structure, al-
with Ba2+ in the tunnels. In similar fashion, coronadite is de- though the weak intensity of the latter. The FT–IR spectrum
fined as hollandite with Pb2+ ions. These oxides crystallise of todorokite (Mn,Ca)Mn3 O7 ·H2 O differs from romanechite
in the tetragonal symmetry (I4/m space group). by three dominant bands located at 414, 493 and 530 cm−1
The RS spectrum of NMD hollandite (Fig. 5, curve a) which are the fingerprint of the T3,3 structure. The different
shows three dominant bands at 628, 586, and 507 cm−1 , shape could be assigned to the nature of the (Ca,Mn)–oxygen
while three weak bands are recorded at 705, 558, and bond lengths perpendicular to the MnO6 octahedral chains
395 cm−1 . Due to the T2,2 tunnel lattice and the double-chain including the presence of Mn3+ ions. Consequently, these
structure, the spectral shape of ␣-MnO2 resembles to that of two effects intensify the low-wavenumber stretching modes
R–MnO2 . The RS spectrum of coronadite PbMn7 O14 dis- at 419 cm−1 . The shift of the high-wavenumber band appear-
plays hollandite-like spectral features which are very weak ing at 748 cm−1 is attributed to the presence of asymmetric
except the band at 585 cm−1 (Fig. 5, curve b). This impor- Mn–O stretch from the triple chains of the todorokite lattice.
tant attenuation of the Raman activity is attributed to the The basic layered structure of the birnessite group (T1,∞ )
presence of heavy ions which damp the MnO6 vibrational is similar to that of chalcophanite [44]. MnO6 octahedral
components perpendicularly to the octahedral chains. layers have a vacancy in one over every six octahedral
The FT–IR spectrum of hollandite displays six bands sites. The interlamellar space contained mono- or di-valent
which appear at higher wavenumber than that of coronadite cations (Li+ , Na+ , Zn2+ or Mn2+ ) more or less hydrated.
(Fig. 6, curve a). However, for coronadite (Fig. 6, curve b) MnO1.85 ·0.6H2 O birnessite supports this layered structure,
the band at 571 cm−1 has lower intensity than the band at in which Mn2+ ions together with water molecules occupy
579 cm−1 of ␣-MnO2 . The stretching mode of MnO6 octa- the interlayer space. Its RS spectrum is dominated by three
hedra along the double-chain is located at 708 cm−1 in the bands at 506, 575, and 646 cm−1 which are attributed to
FT–IR spectra of the hollandite group. In the far-infrared re- the stretching vibrations of MnO6 octahedra (Fig. 5, curve
gion, the band at 310–314 cm−1 is diagnostic of minerals of e). The correlation between the frequency position of the
the hollandite structure. It is noticeable that the far-infrared Raman-active modes and the degree of polymerisation
spectral features involve mainly vibrations of the octahedral shows that MnO1.85 ·0.6H2 O birnessite has an average of
MnO6 skeleton and do not display a significant contribution 4.8 shared edges per MnO6 octahedron [45]. The FT–IR
from the large cations in the channel. spectrum of birnessite (Fig. 6, curve e) displays the open-
ing nature of the compound. As each MnO6 octahedral
chains is separated from another by a distance di = 7 Å, the
Table 6
Infrared modes of mineral manganese dioxides
stretching modes are shifted towards the low-wavenumber
side. Thus, the prominent infrared bands are located at 423,
Compound Infrared wavenumber (cm−1 ) 477, and 513 cm−1 . The sharpness of the band at 423 cm−1
␯1 ␯2 ␯3 ␯4 ␯5 ␯6 ␯7 ␯8 ␯9 indicates the crystalline order of the birnessite compound.
Nsutite 228 330 – 380 472 533 567 690 –
As a concluding remark, it is obvious that the difference in
Hollandite – 314 – 392 463 530 579 710 – the Raman and infrared spectra of the studied minerals here
Coronadite – 310 – – 482 520 571 709 – are due to the identity of the tunnel structure.
Romanechite 275 319 400 432 461 520 577 716 748 The ␣-MnOOH (groutite) crystallises in the orthorhombic
Todorokite 231 – 419 – 493 530 620 – 748 structure in which large variations in Mn–O bond lengths
Birnessite – – 423 – 477 513 583 635 –
are attributed to the presence of the Mn3+ Jahn–Teller ion.
696 C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700

0.6 350
Groutite
510

α-MnOOH
580
Absorbance units

300 α-Mn2O3
616

0.4
1022

(a)
995
360
296

380

RS intensity (arb. units)


329

250
γ-Mn2O3
872

0.2 (b)
200

0.0 150
200 400 600 800 1000 1200 λ-MnO2
(a) Wavenumber (cm-1) (c)

100
100
552

80 50
Groutite Mn3O4
RS intensity (cps)

(d)
60
528

0
384

200 300 400 500 600 700 800 900


142

615

40
Raman shift (cm-1)
352
213
253

648

20 Fig. 8. Raman scattering spectra of manganese oxides: (a) bixbyite


␣-Mn2 O3 , (b) ␥-Mn2 O3 , (c) spinel ␭-MnO2 , and (d) hausmannite Mn3 O4
(IBA100 sample).
0
200 400 600 800 1000
(b) Raman shift (cm-1)
Fig. 7. (a) FT–IR and (b) Raman spectrum of groutite ␣-MnOOH.
(1 × 1) channels. The crystallographic structure of ␭-MnO2
possesses the space group Fd3m (O7h symmetry) and has
a general structural formula A[Mn2 ]O4 , where the Mn4+
The oxygen atoms form distorted octahedra around the man- cations reside on the octahedral (16d) sites, the oxygen
ganese atoms. In this structure, MnO6 octahedra share edges anions occupy the (32e) sites, and A are voids. Analysis of
to form double strings parallel to c-axis [46]. Fig. 7 shows the vibrational spectra of spinel ␭-MnO2 can be made by
the RS and FT–IR spectra of natural groutite. The RS spec- the classical factor-group theory using spectroscopic sym-
trum displays well-resolved and sharp bands with the main metry O7h [48]. The total irreducible representation for the
components at 142, 352, 384, 528, 552 and 615 cm−1 . As the O7h symmetry gives 33 phonons which, for ␭-MnO2 , reduce
elementary cell contains four MnOOH units, 21 Raman- and to four Raman-active modes (A1g + Eg + 2F2g ) and three
15 infrared-active modes are expected from the factor-group infrared-active vibrations (3F1u ). The vibrational modes of
theory. RS bands at 552 and 615 cm−1 are attributed to the Raman and IR allowed phonons result in the loss of
the stretching modes of MnO6 octahedra. The FT–IR spec- F2g and F1u phonons with the removal of lithium from the
trum of natural groutite shows three groups of bands. In the spinel ␭-LiMn2 O4 phase.
far-infrared region bands at 296, 329, 360 and 380 cm−1 are Figs. 8 and 9 (curve c) show the Raman and FT–IR spectra
observed while the stretching vibrations of Mn–O bonds ap- of ␭-MnO2 , respectively. The dominant feature of the Raman
pear at 510, 580 and 616 cm−1 . Additional components at spectrum is the strong band centred at 596 cm−1 attributed
995 and 1022 cm−1 are attributed to Mn–O–H structural vi- to the A1g species. This is the stretching mode of MnO6 oc-
brations in agreement with the spectral features reported by tahedra in ␭-MnO2 . Three weak bands appear at 461, 501,
McBride [47]. and 646 cm−1 . When the spinel ␭-LiMn2 O4 lattice is fully
delithiated, the F2g modes become vibrations of oxygen
ions only. To compare with the vibration frequencies of the
6. Lattice vibration of synthetic manganese oxides spinel LiMn2 O4 , calculations of lattice vibrational modes
of ␭-MnO2 have shown that the wavenumbers of the Eg and
X-ray powder diffraction pattern of ␭-MnO2 displays the F2g modes are expected to increase, while the wavenum-
cubic-close-packed [Mn2 ]O4 spinel framework structure ber of the A1g mode is predicted to decrease [20]. The
(Fig. 1(f)), where the Mn cations are distributed in alter- experimental data for ␭-MnO2 support these predictions.
nate layers in a 3:1 ratio. The interstitial space of ␭-MnO2 The peak positions of the A1g and F2g modes shift toward
contains a three-dimensional network of narrow interlinked the low-frequency side by 27 and 31 cm−1 , respectively.
C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700 697

2.0 A1g mode, which corresponds to the Mn–O breathing vibra-


tion of divalent manganese ions in tetrahedral coordination.
α-Mn2O3 In the vibrational modes of species A1g and Eg , only motions
of the oxygen atoms are involved. Thus the breathing mode
(a) of the tetrahedral MnO6 units is mainly influenced by bond-
1.5 ing and repulsion effects of this unit, e.g. the covalency of
Absorbance (arb. units)

γ-Mn2O3
the Mn–O bonds. Two smaller peaks are located at 300–310
and 350–360 cm−1 . It also has a weak signal at 485 cm−1 .
These RS features are in good agreement with the literature
1.0
(b) data [29]. However, we can notice wide variation in peak
position of the A1g mode of Mn3 O4 from 650 to 668 cm−1 .
λ-MnO2 FT–IR spectra of the synthetic product prepared by thermal
(c) treatment of MnO2 at 1000 ◦ C show similar features to that
of mineral hausmannite.
0.5 The manganese sesquioxide bixbyite appears with two
(d) Mn3O4
polymorphs, the ␣-Mn2 O3 and the tetragonal ␥-Mn2 O3 .
␣-Mn2 O3 bixbyite (C-type structure) has a body-centered
cubic lattice with space group Ia3 (Th7 ). Factor-group
analysis predicts 16 infrared-active modes (16Fu ) and 22
0.0
100 200 300 400 500 600 700 800 Raman-active modes (4Ag + 4Eg + 14Fg ) [27]. Raman and
Wavenumber (cm-1) IR spectra of a Mn2 O3 commercial sample are shown in
Figs. 8 and 9 (curve b), respectively.
Fig. 9. FT–IR powder absorption spectra of manganese oxides: (a) bixbyite
␣-Mn2 O3 , (b) ␥-Mn2 O3 , (c) spinel ␭-MnO2 , and (d) hausmannite Mn3 O4 The IR spectrum divides naturally into three sets of broad
(IBA100 sample). bands (Fig. 9(b)). The high-frequency range presents a sin-
gle band located at 670 cm−1 . The strongest bands in the
spectrum occur in the range of 350–600 cm−1 and appear as
The FT–IR spectrum of ␭-MnO2 (Fig. 9, curve c) displays a broad, rather featureless region of absorption with the max-
broad and intense band centred at 480–520 cm−1 and bands imum absorption near 480 cm−1 . The low-frequency range
with medium intensity at 314, 348, 414 and 603 cm−1 . The 150–300 cm−1 presents an assortment of weak bands. If we
two higher wavenumber modes involve mainly displace- assume that each of the sets must contain at least the num-
ments of oxide ions, although the vibrations are attributed ber of bands observed in the spectrum with best resolution,
to the Mn–O stretching mode, while infrared bands at all 16 bands predicted by the factor-group theory analysis
300–420 cm−1 are assigned to the deformation mode of can be accounted for. The complex IR features of Mn2 O3 is
O–Mn–O bonds. As predicted by lattice dynamics calcula- also due to the structural arrangement of Mn–O bonds. Con-
tions, the mode at 276 cm−1 is absent in the IR spectrum sidering the coordination polyhedra in the C-type structure,
of ␭-MnO2 . the M(1) polyhedron is essentially regular and has a single
The hausmannite Mn3 O4 (MnMn2 O4 in spinel nota- metal–oxygen distance (d1 ), while the M(2) polyhedron is
tion) is a normal tetragonal spinel structure with space quite distorted with three pairs of metal–oxygen distances
group I41 /amd (D19 4h ). The elementary unit cell con- (d2 ) varying considerably in length. For the Mn2 O3 bixbyite,
tents Mn4 Mn8 O16 . Factor-group analysis predicts 10 these distances have been reported to be d1 = 2.01 Å and
infrared-active modes (4Au + 6Eu ) and 14 Raman-active d2 = 1.90–2.24 Å [23]. The broadness of the IR bands is
modes (2A1g + 2B1g + 4B2g + 6Eg ). This compound is attributed to this wide distribution of the metal–oxygen dis-
highly stable under the laser beam. Raman and IR spec- tance. The highest frequency in the IR spectrum is related
tra of hausmannite Mn3 O4 sample are shown in Figs. 8 to the shortest Mn(1)–O distance.
and 9 (curve d). The IR spectrum of Mn3 O4 displays sharp RS spectrum of the black bixbyite is very weak and only
peaks in the range of 100–700 cm−1 . Four bands with six Raman lines could be measured, which is less than half
strong oscillator strength are observed at 241, 396, 475 of the 22 predicted modes. The spectrum of Mn2 O3 bixbyite
and 594 cm−1 . At least the number of bands observed in (Fig. 8(b)) displays two strong Raman bands located at 314
the spectrum with best resolution is higher than the modes and 592 cm−1 and four weak peaks at 192, 481, 645 and
allowed by the factor-group theory analysis. This result is 698 cm−1 . One can remark that powder of Mn2 O3 gives fea-
in good agreement with that reported by Lutz et al. [29]. tures similar to MnO2 . These Raman lines correspond to
The RS spectrum of Mn3 O4 is characterised by a very those reported by Buciuman et al. [26]. They observed the
sharp peak at 654 cm−1 , which was found on mineral haus- influence of the wavelength of the laser excitation on the Ra-
mannite, as well as on either chemically prepared sample man features. Mn2 O3 gave a peak at 640–650 cm−1 , which
or commercial powder. This peak is characteristic of all the was much greater in violet and green light than in red light
spinel structures, illustrated by Fe3 O4 . It is assigned to the and a pair of features at ca. 310 and 360–390 cm−1 . These
698 C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700

authors assume that RS peaks are owing to the formation 1.6


of Mn3 O4 during the spectra acquisition. TGA and DTA 329
MnO
studies show that the transition from pyrolusite to bixbyite
Absorbance units

(Alpha product)
occurs at 500 ◦ C and the one’s bixbyite-hausmannite over 1.2
800 ◦ C. These transitions accompanied the decrease in oxi- 473
dation state of manganese ion from Mn4+ , Mn3+ to Mn2.66+ .
258

0.8
Curiously there is no similarity (see Table 3) between our
Mn2 O3 spectrum and that reported by White and Kerami-
das [27], Kapteijn et al. [22] and Bernard et al. [24]. Our 0.4
concluding remark is that we consider as characteristic for
␣-Mn2 O3 is the peak at 592 cm−1 , the main feature of the
0.0
Mn–O stretching modes of bixbyite. 200 300 400 500 600 700 800 900
Figs. 8 and 9 (curve a) show the RS and FT–IR spectra (a) Wavenumber (cm -1)
of ␥-Mn2 O3 obtained by partial thermal decomposition of
MnO2 . The main Raman feature of ␥-Mn2 O3 is observed at
628 cm−1 with smaller peaks at 263 and 308 cm−1 and very
60
small bands at 383, 598 and 670 cm−1 . The FT–IR spec- MnO
531

Manganosite
trum of ␥-Mn2 O3 is quite similar to that of ␣-Mn2 O3 . Three
RS intensity (cps)

strongest IR bands are located at 446, 484 and 591 cm−1 .


654

40
The Raman features of ␥-Mn2 O3 , reported by Bernard et al.
591
250

[24] show the appearance of a band located at 633 cm−1 in


the case of MnO2 excited with a 20 mW laser power. This
was explained by the formation of a new compound, called 20
XM phase, which was identified as ␥-Mn2 O3 [24]. This
phase has a distorted hausmannite structure, an intermediate
formed during the decomposition of MnO2 to Mn3 O4 . 0
200 300 400 500 600 700 800 900
The mineral manganosite, MnO, has a cubic NaCl-type (b)
structure (space group Fm3m). Fig. 10 shows the vibrational Raman shift (cm -1)
spectra of pelletised MnO. The FT–IR spectrum of MnO Fig. 10. (a) FT–IR and (b) Raman spectrum of manganosite MnO (Alfa
is dominated by a broad band at 329 cm−1 associated with Aesar product).
two shoulders at 258 and 473 cm−1 (Fig. 10(a)). The Raman
spectrum of MnO is characterised by the presence of two (MnO having the NaCl structure) with eight shared edges
broad bands at 531 and 654 cm−1 , while small features ap- per MnO6 octahedron that gave the lowest wavenumber
pear at 250 and 591 cm−1 (Fig. 10(b)). The strongest peak ν1 = 531 cm−1 . Thus, similar to the IR features (Fig. 11),
located at 531 cm−1 is attributed to the Mn–O stretching the graph presented in Fig. 12 suggests that Raman spectra
mode of the Mn2+ cations in the cubic cage. Similar features
have been observed by Bernard et al. [24] who reported RS
bands at 521, 595, and 645 cm−1 . Pyrolusite Ramsdellite
Romanechite
Manganosite

800
7. Vibrational features versus MnO6 polymerisation
ν1
Raman shift (cm -1)

700
Taking into account the correlation of infrared spectra
made by Potter and Rossman [41], the spectral identifica- 600
tion of the polymerisation by edge-sharing of the MnO6 ν2
octahedra in manganese oxides was obtained from Raman
500
data. Figs. 11 and 12 present the frequency position of the
major infrared- and Raman-active modes, respectively, as
MnO2
a function of the average MnO6 octahedral polymerisa- 400
compounds
tion in several manganese oxide structures. A trend is also
observed showing a general decrease in band wavenum- 300
ber with increasing octahedral polymerisation. The two 0 2 4 6 8
sets of stretching modes, noted ␯1 and ␯2 , allow a direct Edges shared per MnO 6 octahedron
comparison between the Raman patterns of manganese ox- Fig. 11. Frequency position of the Raman-active modes as a function of
ides. Starting from pyrolusite with two shared edges per the average MnO6 octahedral polymerisation in manganese oxides. Dot
MnO6 octahedron, one reaches the point for manganosite size is approximately proportional to band intensity.
C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700 699

Ramsdellite 8. Conclusion
Pyrolusite Manganosite
Romanechite
A systematic study of vibrational features has been con-
800 ducted in order to compare the structural and morphological
properties of various manganese oxide lattices. As Raman
700
Band position (cm-1)

spectroscopy is not a primary structural technique like


600 X-ray diffraction, it was necessary to calibrate it against
well-crystallised materials. Experimental features have
500 shown that Raman and infrared spectroscopy are sensitive
tools to probe the short-range environment of oxygen co-
400 ordination around transition-metal cation in oxide lattices.
300 The differences in the vibrational behaviour of MDOs have
been mainly attributed to the lattice polymerisation, i.e. the
200 edge-sharing of the MnO6 octahedra, and to the intrinsic
0 2 4 6 8
Edges shared per MnO6 octahedron distortion of the basic MnO6 octahedron but also in some
extends to the variations in average manganese oxidation
Fig. 12. Frequency position of the infrared-active modes as a function of states. Our results confirm the spectral features of various
the average MnO6 octahedral polymerisation in manganese oxides. Dot manganese oxides displaying bands which are diagnostic
size is approximately proportional to band intensity. of their Mn–O skeleton and structural arrangement of basic
MnO6 entities.
are dependent on the MnO6 octahedral framework rather
than oxidation state of cations. To apply this correlation,
Acknowledgements
the position of the low-energy Raman band of the bir-
nessite MnO1.86 ·0.6H2 O (Fig. 6) supports its proposed
The authors thanks Mr. M. Lemal for his technical assis-
layered structure, which places it at 4.8 shared edges per
tance in XRD measurements. Dr. J.C. Bouillard (Collection
octahedron.
de Minéralogie, Université Pierre et Marie Curie), Dr. D.
Fig. 13 presents the variations of the Raman stretching
Guyomard (IMJR, Nantes), Dr. B. Yebka (GM R&D Cen-
band position with the average Mn–O distance in man-
tre, Warren), and Dr. J.P. Pereira-Ramos (CNRS, Thiais) are
ganese dioxide lattices. This graph shows that a correlation
gratefully acknowledged for providing mineral and synthetic
exists between Mn–O bond lengths and stretching frequen-
manganese dioxide samples.
cies, which can be held relatively well by a second order
polynomial fit. The shortest Mn–O bond, 1.86 Å, is as-
signed to the highest occurring stretching Raman frequency
References
at 750 cm−1 for the octahedrally coordinated MnO6 groups.
As early demonstrated by Hardcastle and Wachs [49] for
[1] Q. Feng, H. Kanoh, K. Ooi, J. Mater. Chem. 9 (1999) 319.
vanadium oxides, this empirical correlation could lead [2] P. Le Goff, N. Baffier, S. Bach, J.P. Pereira-Ramos, Mater. Res. Bull.
to a systematic determination for the spectral features of 31 (1996) 63.
manganates. [3] J.C. Hunter, J. Solid State Chem. 39 (1981) 142.
[4] M.M. Thackeray, A. de Kock, L.A. de Piocciotto, G. Pistoia, J.
Power Sources 26 (1989) 355.
[5] J.C. Nardi, J. Electrochem. Soc. 132 (1985) 1787.
[6] C. Palache, H. Berman, C. Frondel, The System of Mineralogy, 7th
Raman wavenumber (cm-1)

750 MDO compounds ed., vol. 1, Wiley, New York, 1963.


[7] T. Ohzuku, M. Kitagawa, T. Hirai, J. Electrochem. Soc. 137 (1989)
3169.
[8] P. Le Goff, N. Baffier, S. Bach, J.P. Pereira-Ramos, J. Mater. Chem.
700 4 (1994) 875.
[9] C. Julien, M. Massot, C. Perez-Vicente, E. Haro-Poniatowski, G.A.
Nazri, A. Rougier, Mater. Res. Soc. Symp. Proc. 496 (1998) 415.
[10] C. Julien, A. Rougier, E. Haro-Poniatowski, G.A. Nazri, Mol. Cryst.
Liq. Cryst. 311 (1998) 81.
650
[11] C. Julien, A. Rougier, G.A. Nazri, Mater. Res. Soc. Symp. Proc. 453
(1997) 647.
[12] W. Huang, R. Frech, J. Power Sources 81–82 (1999) 616.
1.8 1.9 2.0 2.1 2.2 [13] C. Julien, Ionics 6 (2000) 30.
Mn-O average distance (Å) [14] C. Julien, in: C. Julien, Z. Stoynov (Eds.), Materials for Lithium-ion
Batteries, NATO Science Series, vol. 3–85, Kluwer Academic Pub-
Fig. 13. Variations of the Raman stretching band position with the average lication, Dordrecht, 2000, p. 309.
Mn–O distance in manganese dioxide lattices. [15] C. Julien, Solid State Ion. 136–137 (2000) 887.
700 C.M. Julien et al. / Spectrochimica Acta Part A 60 (2004) 689–700

[16] C. Julien, S. Rangan, M. Lemal, M. Massot, D. Guyomard, J. Raman [32] M. Mermoux, C. Poinsignon, P.O. Robert, ITE Lett. 2 (2001) 59.
Spectrosc. 33 (2002) 223. [33] J.B. Fernandes, B. Desai, V.N. Kamat Dalal, Electrochim. Acta 28
[17] T.J. Richardson, P.N. Ross Jr., Mater. Res. Bull. 31 (1996) 935. (1983) 309.
[18] T.J. Richardson, S.J. Wen, K.A. Striebel, P.N. Ross Jr., E.J. Cairns, [34] G. Gattow, O. Glemser, Z. Anorg. Allg. Chem. 309 (1961) 121.
Mater. Res. Bull. 32 (1997) 609. [35] O. Glemser, G. Gattow, H. Meisiek, Z. Anorg. Allg. Chem. 309
[19] S.J. Wen, T.J. Richardson, L. Ma, K.A. Striebel, P.N. Ross Jr., E.J. (1961) 1.
Cairns, J. Electrochem. Soc. 143 (1996) L136. [36] G.A. Kolta, F.M. Abdel Kerim, A.A. Abdul Azim, Z. Anorg. Allg.
[20] B. Ammundsen, G.R. Burns, M.S. Islam, H. Kanoh, J. Roziere, J. Chem. 384 (1971) 260.
Phys. Chem. B 103 (1999) 5175. [37] M. Ishii, M. Nakahira, T. Yamanaka, Solid State Commun. 11 (1972)
[21] B.R. Strohmeier, D.M. Hercules, J. Phys. Chem. 88 (1984) 4923. 209.
[22] F. Kapteijn, A.D. van Langeveld, J.A. Moulijn, A. Andreini, M.A. [38] F. Fillaux, C.H. Cachet, H. Ouboumour, J. Tomkinson, C.
Vuurman, A.M. Turek, J.M. Jehng, I.E. Washs, J. Catal. 150 (1994) Levy-Clement, L.T. Yu, J. Electrochem. Soc. 140 (1993) 585.
94. [39] D.A.J. Swinkels, K.E. Anthony, P.M. Fredericks, P.R. Osborn, J.
[23] D. Gosztola, M.J. Weaver, J. Electranal. Chem. Interface Elec- Electroanal. Chem. 168 (1984) 433.
trochem. 271 (1989) 141. [40] J.M. Amarilla, L.A.H. MacLean, F. Tedjar, F. Le Cras, P. Strobel,
[24] M.C. Bernard, A. Hugot-Le Goff, B.V. Thi, S. Cordoba de Torresi, C. Poinsignon, Mater. Res. Soc. Symp. Proc. 369 (1995) 87.
J. Electrochem. Soc. 140 (1993) 3065. [41] R.M. Potter, G.R. Rossman, Am. Mineral. 64 (1979) 1199.
[25] B.A. Lopez de Mishima, T. Ohtsuka, N. Sato, J. Electroanal. Chem. [42] L.A.H. McLean, C. Poinsignon, J.M. Amarilla, F. Le Cras, P. Strobel,
243 (1988) 219. J. Mater. Chem. 5 (1985) 1183.
[26] F. Buciuman, F. Patcas, R. Craciun, D.R.T. Zahn, Phys. Chem. Chem. [43] L.A.H. McLean, C. Poinsignon, ITE Lett. Batt. 1 (2000) 87.
Phys. 1 (1999) 185. [44] R. Giovanoli, E. Stahli, W. Feitknecht, Helv. Chem. Acta 53 (1970)
[27] W.B. White, V.G. Keramidas, Spectrochim. Acta 28A (1972) 501. 453.
[28] H.H. Chou, H.Y. Fan, Phys. Rev. B 13 (1976) 3924. [45] A. Manceau, V.A. Drits, E. Silvester, C. Bartoli, B. Lanson, Am.
[29] H.D. Lutz, B. Muller, H.J. Steiner, J. Solid State Chem. 90 (1991) Mineral. 82 (1997) 1150.
54. [46] L.S.D. Glasser, L. Ingram, Acta Crystallogr. B 24 (1968) 1233.
[30] V. Maroni, J. Phys. Chem. Solids 49 (1988) 307. [47] M.B. McBride, Soil Sci. Soc. Am. 51 (1987) 1466.
[31] R.G. Buckley, B.P. Clayman, J.C. Mikkelsen Jr., Phys. Rev. B 26 [48] G.C. Allen, M. Paul, Appl. Spectrosc. 49 (1995) 451.
(1982) 6509. [49] F.D. Hardcastle, I.E. Wachs, J. Phys. Chem. 95 (1991) 5031.

You might also like